Next Article in Journal
The Influence of Different Operation Conditions on the Treatment of Mariculture Wastewater by the Combined System of Anoxic Filter and Membrane Bioreactor
Next Article in Special Issue
Efficient Photocatalytic Degradation of Organic Pollutant in Wastewater by Electrospun Functionally Modified Polyacrylonitrile Nanofibers Membrane Anchoring TiO2 Nanostructured
Previous Article in Journal
A Novel Numerical Procedure to Estimate the Electric Charge in the Pore from Filtration of Single-Salt Solutions
Previous Article in Special Issue
Experimental Study on a Ceramic Membrane Condenser with Air Medium for Water and Waste Heat Recovery from Flue Gas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ionic Liquid in Phosphoric Acid-Doped Polybenzimidazole (PA-PBI) as Electrolyte Membranes for PEM Fuel Cells: A Review

by
Leong Kok Seng
1,2,
Mohd Shahbudin Masdar
1,3,* and
Loh Kee Shyuan
3
1
Department of Chemical and Process Engineering, Universiti Kebangsaan Malaysia, Bangi 43600, Selangor, Malaysia
2
Department of Petrochemical Engineering, Politeknik Tun Syed Nasir Syed Ismail, Pagoh 84600, Johor, Malaysia
3
Fuel Cell Institute, Universiti Kebangsaan Malaysia, Bangi 43600, Selangor, Malaysia
*
Author to whom correspondence should be addressed.
Membranes 2021, 11(10), 728; https://doi.org/10.3390/membranes11100728
Submission received: 28 July 2021 / Revised: 17 September 2021 / Accepted: 20 September 2021 / Published: 24 September 2021

Abstract

:
Increasing world energy demand and the rapid depletion of fossil fuels has initiated explorations for sustainable and green energy sources. High-temperature polymer electrolyte membrane fuel cells (HT-PEMFCs) are viewed as promising materials in fuel cell technology due to several advantages, namely improved kinetic of both electrodes, higher tolerance for carbon monoxide (CO) and low crossover and wastage. Recent technology developments showed phosphoric acid-doped polybenzimidazole (PA-PBI) membranes most suitable for the production of polymer electrolyte membrane fuel cells (PEMFCs). However, drawbacks caused by leaching and condensation on the phosphate groups hindered the application of the PA-PBI membranes. By phosphate anion adsorption on Pt catalyst layers, a higher volume of liquid phosphoric acid on the electrolyte–electrode interface and within the electrodes inhibits or even stops gas movement and impedes electron reactions as the phosphoric acid level grows. Therefore, doping techniques have been extensively explored, and recently ionic liquids (ILs) were introduced as new doping materials to prepare the PA-PBI membranes. Hence, this paper provides a review on the use of ionic liquid material in PA-PBI membranes for HT-PEMFC applications. The effect of the ionic liquid preparation technique on PA-PBI membranes will be highlighted and discussed on the basis of its characterization and performance in HT-PEMFC applications.

1. Introduction

Rising energy demand driven by the growth of community and industries, catalyzed by the rapid depletion of petroleum-based energy sources, has initiated the interest of academicians or researchers to find new energy sources. Furthermore, energy generated from petroleum-based sources produces hazardous gases such as CO, NOx and SOx that are harmful to the environment, initiating acidic rain, depletion of stratospheric ozone and global climate change. Thus, a new type of energy producer is necessary to meet the demands while being environmentally friendly.
To date, electrical energy is the most used energy, and demands for this energy are rising with population growth. Approximately 86.4% of electrical energy is from fossil fuels, such as oil (36%), natural gas (23%) and coal (27.4%), with the remaining 14.6% generated from renewable sources [1,2,3,4]. Thus, the concept of fuel cells has emerged as an innovative energy conversion device that is considered a sustainable and environmentally friendly energy conversion procedure. Fuel cells operate via an electrochemical reaction directly converting chemical energy into electrical energy (Figure 1). This technology is considered the most promising power generation and offers several advantages including high efficiency, reliability, zero-emission, silent operation and low maintenance [5,6,7]. There are five types of fuel cells classified based on the electrolytes, namely polymer electrolyte membrane fuel cell (PEMFC), alkaline fuel cell (AFC), phosphoric acid fuel cell (PAFC), molten carbonate fuel cell (MCFC) and solid oxide fuel cell (SOFC).
PEMFC has emerged as the most promising fuel cell technology due to its high energy efficiency and power density, low emissions, compact size, light weight, fast start-up, short refueling time and low operating temperature [9,10]. There are two types of fuel cells, low- or high-temperature PEMFC, classified based on their operating temperatures. Low-temperature (LT) PEMFC usually operates below 100 °C and demonstrates exceptional performance with a maximum power density of about 500 to 100 mW cm−2 under H2/air [11,12]. However, the larger radiator volume and complex water management hinder the application of these fuel cells. In contrast, the high-temperature (HT) PEMFC operates from 100 to 200 °C, and the relatively simple design requires a simple radiator so external humidification is not necessary [13,14]. HT-PEMFC also has several advantages, such as increased electrode kinetics, higher tolerance to CO, low crossover and few by-products [15,16].
For the development of PEMFC, the most significant factor is a highly proficient polymer electrolyte membrane. Initially, perfluorosulfonic acid polymer membranes such as Nafion were used, as these materials demonstrated good conductivity, chemical and mechanical stability, as well as a higher power density [17,18]. However, this polymer fails to perform at high operating temperatures due to decreased proton conductivity and destruction of the polymer structure [19,20]. Thus, a new polymer, polybenzimidazole (PBI), is a promising alternative due to its chemical and thermal stability without humidification and low cost [21]. These polymers also contain amide or imide groups that can function as proton acceptors and react in an acidic medium [22].
To increase the conductivity of the PBI membrane, several mineral acids can be used as dopants, such as HNO3, H2SO4, HClO4 and HCl. Xing and Savadogo [23] proposed the conductivity order of mineral acids as follows H2SO4 > H3PO4 > HCIO4 > HNO3 > HCl. Even though sulfuric acid ranked higher in the list, it is not practical to apply for doped acid as more than 50% relative humidity (RH) is needed to obtain the maximum output of >0.2 S/cm at 150 °C with a doping level of 9.65 [24]. Moreover, the stability of the PBI membrane decreased rapidly in hot concentrated sulfuric acid. Thus, phosphoric acid is frequently selected as a dopant due to its higher conductivity, outstanding thermal stability and low vapor pressure at high temperatures [25]. Jones and Rozière [26] also explained that the presence of free acids in the polymer structure and H2PO4/HPO42− anionic chains initiated higher proton conductivity of the PA-PBI polymer.
The conductivity of the PA-PBI polymer depends on the amount of phosphoric acid-doped in the polymer. However, a higher amount of phosphoric acid leads to the degradation of mechanical properties and acid leaching, thus limiting the conductivity of the pure PA-PBI. The most efficient method to overcome this problem is introducing ionic liquids into the polymer phase [9].

2. PBI Membrane in PEMFC

In the last decade, considerable efforts have been made to develop high-temperature (>100 °C) PEMFCs using polymer acid complexes (PACs), as they offer significant advantages in this temperature range, such as (1) improved CO tolerance, (2) enhanced efficiency, (3) avoidance of flooding by-water, (4) opportunity to use non-noble metal catalysts and (5) system simplification. Early investigation of PEMFC applied structure Nafion membrane (Figure 2) as a proton exchange membrane, and to date this material has been recognized as a reference for PEMFC [27]. This perfluorinated type of membrane was commercialized by the DuPont company and has several significant characteristics, such as high proton conductivity as well as good chemical and mechanical properties, for fuel cell operations with more than 60,000 h of operation.
However, Nafion has several drawbacks that affect the performance of fuel cells. This membrane must be firstly hydrated because the conductivity and productivity decrease quickly with decreasing RH, high methanol crossover and costly materials. Moreover, the complex design including water and heat management must be built in, as the operation of this fuel cell requires sufficient water content in the membrane to preserve the membrane conductivity and maintain the operating temperature below 80 °C.
The requirements for high-performance PEMFCs are as follows:
(1)
Proven stability in terms of electrochemical, chemical and thermal for fuel cell operations;
(2)
Higher mechanical tensile strength and sturdiness under heavy loads;
(3)
Higher gas separation capacity;
(4)
Good electrical insulation;
(5)
Cost-effective.
The material membranes selected have a direct impact on the storage device’s performance in a wide range of applications. Fuel cells are a promising technique to produce an environmentally friendly conversion energy system stored in a fuel. A hydrogen economy based on renewables, which includes hydrogen production, storage and power conversion, has been widely seen as a potential answer for the future of energy. Hydrogel electrolytes that are alkali-tolerant have been widely used in next-generation alkaline energy devices [28]. Preoxidized kraft lignin and poly(ethylene glycol)diglycidyether (PEGDGE) crosslinking reactions have been constructed, studied and used as quasi solid-state (QS) electrolytes in aqueous dye-sensitized solar cell (DSSC) devices, which showed a straightforward strategy for the field of sustainable photovoltaic devices [29]. The core double shell photocatalyst was a promising, magnetically separable and stable photocatalyst for long-term practical applications of photo oxidation [30]. For energy storage devices, a nanocomposite of CoSn alloy with a multishell layer structure enclosed in 3D porous carbon showed excellent performance when used as an anode for lithium-ion batteries (LIBS) [31]. A composite gel polymer electrolyte consisting of a highly cross-linked polymer matrix, containing a dextrin-based nanosponge and activated with a liquid electrolyte, exhibited good ionic conductivity at room temperature [32]. Lithium bis(trifluoromethylsulfonyl)imide (LiTFSI) on a solid polymer electrolyte (SPE) system with 30 wt.% LiTFSI doping level achieved an ionic conductivity of 3.69 × 10−8 Scm−1 at ambient temperature and 1.23 × 10−4 Scm−1 at 373 K [33,34].
Since the first successful application of poly-benzimidazole (PBI) membranes as electrolytes, PBI has been extensively explored as a proton conducting electrolyte in fuel cell applications due to its high thermal, chemical and mechanical stability and high proton conductivity [35]. PBI membranes are also highly resistant to acidic or basic conditions and have high glass transition temperatures (425–436 °C), low flammability, high energy radiation resistance and are relatively inexpensive. The most convenient characteristic of the PBI membrane is that this polymer is suitable for the high temperatures of fuel cells, as acids or alkaline groups act as a proton carrier without hydration.
PBI monomers are linear aromatic heterocyclic macromolecules and were first introduced by Vogel and Marvel in 1961 [36] for defense and aerospace applications. Wainright et al. [37] introduced phosphoric acid-doped PBI membranes as a polymer electrolyte for HT-PEMFCs in 1995, thus initiating research in this area. Several methods were introduced to prepare this membrane, including polymerization in polyphosphoric acid (PPA), casting from methane sulfonic acid and microwave-assisted organic synthesis. To date, poly[2,2′-(m-phenylene)-5,5′-bibenzimidazole], also known as m-PBI, and poly(2,5-benzimidazole) or AB-PBI are the most common PBI membranes used for the study of HT-PEMFC [8]. M-PBI can be prepared via polycondensation of diaminobenzidine monomer with isophthalic acid, while AB-PBI is prepared via polycondensation of 3,4-aminobenzoic acid (DABA) either in PPA or Eaton’s reagent [3,37,38]. Figure 3a shows the chemical structure of the m-PBI and AB-PBI monomers, while Figure 3b,c depicts their synthesis. Figure 3d presents the synthesis of the linear and cross-linking sulfuric acid-OPBI polymer [39], and Figure 4 indicates the possible proton transfer pathway for the sulfuric acid–PBI polymer [40]. Table 1 lists examples of the preparation methods for PBI polymers and their applications.
Mekhilef et al. [50] and Zeis [51] explained that the development of HT-PEMFC was significantly influenced by PAFCs using phosphoric acid as an electrolyte. Theoretically, the proton conductivity of PBI is minimal; thus, the incorporation of secondary proton conducting materials is crucial to support ion conductivity [52]. Therefore, doping with phosphoric acid is commonly used to increase the conductivity of PBI, known as phosphoric acid-PBI or PA-PBI. This breakthrough technology was first reported by Samms et al. [53] in their study on the proton conductivity of PA-PBI via solid-state NMR, which suggested that the mobility of phosphoric acid in the PBI polymer is lower than that of free phosphoric acid. This phenomenon increased the proton conductivity of PBI polymer and maintained high thermal stability without external gas humidification. Melchior et al. [54] also added a selection of PA based on their high proton conductivity and low vapor pressure. Zeis [42] explained the mechanism of proton transfers in phosphoric acid-doped PBI membrane as shown in Figure 5.
Nevertheless, a significant drawback of PA-PBI is that high conductivity depends on the percentage of phosphoric acid loading, with a high concentration of phosphoric acid affecting the mechanical strength of PBI. In high RH conditions, leaching of PA occurs during fuel cell shutdown caused by water condensation, thus reducing the conductivity [55]. Hu et al. [56] reported that the phosphate anions could be adsorbed onto the surface of platinum, which acts as a catalyst in the fuel cell, blocking the active sites, thus causing deactivation of the platinum. Moreover, Asensio et al. [14] and Yu et al. [57] reported that the pyrolysis of PA occurred at 190 °C, thus initiating loss of proton conductivity. A high phosphoric acid content initiated several problems such as reduced mechanical strength, elution of electrolytes, corrosion of the catalyst, leaching and condensation of phosphate groups at high temperature, and formation of oligomers such as pyrophosphoric acid [3,58,59,60,61]. Extensive research has been conducted to overcome this problem with several solutions proposed, such as the introduction of silica and clay [62,63], metal carborane and metal oxides [64], phosphate salts [8], heteropolyacids [65], metal-organic frameworks [66], graphene oxide [67,68] and ionic liquids [69,70,71] (Table 2).

3. Ionic Liquids

In general, the term ionic liquids refers to any ionic form of liquid that has a boiling temperature below the average boiling water temperature and is liquid at ambient temperature [80,81,82,83,84], for example fused salt, molten salt and liquid organic salt. In addition, ionic liquids are non-volatile, non-flammable and exhibit good chemical and thermal stability as well as high ionic conductivity [85,86]. The specific characteristic of ionic liquids is the high ionic conductivity due to high ion density and viscosity, since they contain ions only and are 10–1000 times more viscous than water [87]. Ionic liquids have attracted considerable attention due to their new and tunable physicochemical properties, especially in electrochemical applications [40,81,82].
Ionic liquids can generally be classified into three classes, i.e., aprotic, protic and zwitterionic (Figure 6a), with each class synthesized for a specific application. Aprotic ionic liquids are a mixture of large organic cations such as pyridinium, imidazolium or phosphonium with smaller anions such as bromine, chloride, sulfate and hexafluorophosphate for inorganic anions, or bis(trifluoromethyl sulfonyl)imide for organic anions [88]. Generally, they are synthesized by alkynation of quaternization of amine groups, followed by an anion exchange reaction (Figure 5b). Aprotic liquids have no active protons in their chemical structure [89]. Protic ionic liquids are synthesized via proton transfer through a Brönsted acid and base, which act as a proton-donor and acceptor and contain exchangeable protons in their chemical structure for hydrogen bond formation (Figure 5b) [90,91]. Zwitterionic ionic liquids are prepared by adding ionic liquid compounds to surfactant systems to modify the surfactant properties (Figure 6c) [85]. Figure 7 shows the most common cations and anions for ionic liquids compounds widely used in the literature, and Figure 8 depicts several types of protic ionic liquid.
Ionic liquids can be prepared for their specific application, as both anions and cations can be incorporated; therefore, they can be used for catalysis, biocatalysis, synthetic chemistry and electrochemistry. Vekariya et al. [95] listed three generations of ionic liquids based on their applications. The first generation of ionic liquids involved the preparation of 1-alkyl-3-methylimidazolium salts by Wilkes et al. in 1982, known as tetrachloroaluminates [96]. Then, the second generation of ionic liquids was successfully developed by replacing the tetrafluoroborate ion and other anions to produce air- and water-stable ionic liquids [97], widely used as solvents for organic reactions. The third generation of ionic liquids was introduced by Davis in 2004 [98], known as task-force ionic liquid, incorporating a large group of cations, such as phosphonium, imidazolium, ammonium, pyridinium and highly diffuse anions, such as BF4, PF6 and CF3SO3. These ionic liquids were synthesized specifically for their applications, for example, acidic chloroaluminate salts contained imidazolium and pyridinium cations for battery applications. Some reports on the usage of ionic liquids are provided in Table 3.
Due to the high conductivity and insignificant vapor pressure of the ionic liquids, these materials are stable to be employed as an additive in mid- and high-temperature PEMFC [9]. Protic ionic liquids containing N and H atoms can form a hydrogen bond network initiating the Grotthuss mechanism of conductivity, which is superior to the vehicle mechanism. Therefore, the addition of a protic ionic liquid enhanced the conductivity of the polymer electrolyte and reduced the inorganic acid dependency, such as phosphoric acid.

4. Ionic Liquids in PBI Membranes

This review is focused on the effects of the ionic liquid doping techniques on the performance of PA-PBI membranes as HT-PEMFCs. The latest techniques and materials applied in ionic liquid doping PA-PBI are also discussed.
As reported by previous studies, the conductivity of the PBI membranes depended on the concentration of phosphoric acid. However, high concentrations of PA are highly corrosive, and this initiated major problems that included damaging the mechanical structure of the cells. Additionally, at higher temperatures, hydration of phosphoric acid and formation of pyrophosphoric acid oligomers reduced the conductivity of PA/PBI [102,103,104]. Therefore, several approaches were proposed, and recently, ILs were presented as a promising solution. ILs contain proton donors and acceptors in their chemical structures, which were expected to enhance the conductivity of the PBI monomers even in low PA concentrations [89].

4.1. Synthesis

Skorikova et al. [108] successfully developed bis(triflioromethanesulfonyl)imide-PBI membranes through direct blending to form quasi-solidified ionic liquid membranes (QSILMs). This approach was recommended for the immobilization of protic ionic liquid compounds in the polymer matrices. Immobilization of protic ILs using this technique offered several advantages such as simple procedure, low consumption of toxic organic solvents and high volume of ILs immobilization. De Trindade et al. [109] also explored the ability of ILs to improve the conductivity of the PBI membranes. They synthesized and characterized the PBI with 3-triethylammonium hydrogen sulfate (TEA) and 1-butylimidazole hydrogen sulfate as the ionic liquid compounds.
Javanbakht et al. [110] employed 1,3-di(3-methylimidazolium) to propane dibromide dicationic ionic liquid (pr(mim)2Br2) as the doping agent for the PA-PBI membranes. The dicationic ILs were classified as ILs as they contained two mono anions and two aromatic rings linked by alkyl chains as cations. The compound had several advantages such as higher thermal stability, glass transition temperature, melting point and proton conductivity compared to mono cationic ILs that improved the quality of the membranes for applications in HT-PEMFCs [111,112]. The investigation used melamine-based dendrimer functionalized-Santa Barbara amorphous-15 mesoporous silica (MDA-SBA-15) as the hydrophilic inorganic particles and exhibited a momentous role in the protection of the PA and pr(mim)2Br2 against water vapor, which was the by-product at the cathode after a long operation time of the HT-PEMFCs.
In another study, Compañ et al. [69] prepared a series of PA/PBI membranes that engaged different exchangeable anions in the ionic liquid to evaluate the effects of the anions and temperature on the proton conductivity of the phosphoric acid-doped PBI membranes. The study applied 1-butyl-3-methylimidazolium (BMIM) as the ionic liquid compound with several anions changed, namely chloride (Cl), bromide (Br), iodide (I), thiocyanate (NCS), bis(trifluoromethylsulfonyl)imide (NTf2), hexafluorophosphate (PF6) and tetrafluoroborate (BF4) ions. The composite membranes were prepared via the casting method with 5 weight percentage (wt.%) of the ILs.
Liu et al. [113] successfully prepared a series of highly conductive cross-linked membranes with fluorine-containing polybenzimidoles (6FPBI) and 1-vinyl-3-butylimidazolium chloride base to form poly(ionic liquid) (PIL) through in situ free radical polymerization. The (PIL) technique was introduced to overcome the leaching of ionic liquid molecules problem while preserving the proton transfer pathway. The PIL was a series of repeating monomers bonded with ionic liquid molecules, either anionic or cationic species of ILs [114,115]. Liu et al. [116] also prepared a series of cross-linked fluorine-polybenzimidazole (6FPBI) membranes with the addition of a cross-linked polymeric ionic liquid for HT-PEM applications. Instead of using linear ionic liquid compounds, they suggested the application of cross-linked polymeric ionic liquid compounds, as the polymeric ILs had several advantages such as providing a continuous and fast pathway for proton transfer with the aids of anions in the polymeric ILs that generally act as proton acceptors. Additionally, multiple proton transport channels were created with the incorporation of polymeric ILs in the polymer matrix, thus preventing the leakage of the ionic liquid compounds.
In a recent study, Gao et al. [117] researched the preparation and characterization of a series of PBI with hyperbranched cross-linked membranes with imidazolium groups that acted as the ionic liquid compounds. Generally, branched polymers showed significant advantages such as good oxidative stability and adsorbed more PA compared to linear polymers. However, the loss of mechanical properties was also observed in the polymers [118]. Thus, the cross-linking method was a promising solution for the improvement of branched polymers’ mechanical strength, as the cross-linkable compounds toughen the interactions among the polymer chains. Additionally, selection of the imidazolium group was crucial, as the group had a conjugated ring structure that accommodated more space for phosphoric acid in the polymer chains, and the delocalization and formation of hydrogen bonds stabilized the adsorbed phosphoric acid, which prevented leakages [119].

4.2. Effect of Ionic Liquids on PA/PBI Membrane Performance

Based on the findings by Liu et al. [112], the formation of PIL was vital to promote proton transfer and to improve the mechanical properties of the PA-PBI membranes. The incorporation of PIL initiated better proton conductivity, more than 76.9% increment at 170 °C, compared to the pristine 6FPBI membranes. Moreover, phosphoric acid’s stability was increased by about 73.1% at 160 °C operating temperature. Epoxy groups in the PIL played a significant role in PA retention as the groups acted as cross-linkers and formed cross-linked networks and prevented leaking of the PA. The increased stability of the PA was expected due to the incorporation of the PIL with dihydrogen phosphate ion (H2PO4). Figure 9 illustrates the synthetic process of 6FPBI and 6FPBI-PIL membranes.
The most significant observation from a study by Liu et al. [115] was that PA retention for this membrane was improved, which prolonged conductivity and stability, even at a longer time of PA doping. The cross-linked membranes also displayed better chemical and oxidative stability and good mechanical properties compared to linear membranes. Moreover, extremely high PA doping levels were achieved, therefore increasing the ionic conductivity of the membranes without any leaking detected. At 170 °C, the proton conductivity of the 6FPBI-cPIL reached about 0.106 S/cm with a doping level of 27.8. Figure 10 depicts the schematic process for the synthesis of (a) cross-linkable polymeric ionic liquid compounds and the (b) preparation of 6FPBI-cPIL. A cross-linkable polymeric liquid compound for this study was prepared via the free radical polymerization of 1-vinyl-3-butylimidazolium trifluoromethanesulfonyl imide ([ViBuIm][TFSI]) with allyl glycidyl ether.
The conductivity of QSILMs prepared by Skorikova et al. [108] achieved about 30–60 mS cm−1 at 180 °C after doping with phosphoric acid compared to the zero IL-PBI membranes, which only achieved less than 10 mS cm−1. This research also suggested a 1:1 ratio of bis(triflioromethanesulfonyl) imide-PBI membrane is the best performing ionic liquid PBI membrane, as this polymer reached a power density of about 0.32 Wcm−2 at 200 °C and 900 mAcm−2. Therefore, ionic liquid compounds have an important role, especially by increasing the conductivity of the PA-PBI membrane by improving retention of doped PA in the membrane with no leaking. The impregnated catalyst layer of the gas diffusion electrode with protic ionic liquid exhibited better stability for long-term use (100 h of operation at 200 °C) compared to phosphoric acid alone. Moreover, this research also suggested the application of fluorescence microscopy for the structural investigation of the PBI membrane and ionic liquid distribution. Generally, PBI contained fluorescence-active molecules in a broad wavelength range [120,121]; thus, the application of fluorescence microscopy will facilitate the analysis of PBI film morphology.
The doping of ionic liquid compounds initiated higher oxidative stability and proton conductivity compared to the non-doped PBI membranes even at higher temperatures and percentages of relative humidity (%RH). TEA ILs produced higher conductivity and oxidative stability compared to BIm ILs, which was caused by the presence of SO3H groups at the cations and SO4H groups at the anions. The H+ generated at the anions and cations increased the number of protons at the membranes, which resulted in improved conductivity compared to the BIm compounds [109]. Javanbakht et al. [110] explored the application of dicationic ILs in the preparation of PA/PBI membranes. From the results of the study, the prepared poly(2,5-benzimidazole)-dicationic ionic liquid (ABPBI-DIL4) showed higher proton conductivity and easier proton exchange, as a lower voltage drop was observed compared to the non-doped ABPBI membranes at a high operating temperature of 180 °C. The observation proved the ability of the ILs to increase the proton conductivity and initiate oxidative stability of the PBI membranes. The oxidative stability was increased with increased PBI content, and the higher conductivity of the membranes was parallel with the increased ILs percentage. Moreover, the formation of hydrogen bonds between acid protons of the ILs cations with the amine group of the PBI prevented the leaching of the ILs.
Compañ et al. [69] observed that the application of ILs as fillers improved the mechanical properties of the PBI membranes, which were caused by the interaction of the polymer matrix and the ionic liquid compounds. The casting technique successfully produced membranes with better thermal, mechanical and oxidative stability, which made the membranes suitable for fuel cell applications. The PBI that contained BMIM-BF4 achieved 94 mS/cm conductivity at 200 °C compared to non-doped PBI membranes, which were observed at about 0.71 mS/cm. Higher conductivity might be initiated by the formation of hydrogen bond networks between the ionic liquid compounds and the PA molecules which were distributed in the polymer matrix. Additionally, the study found that polarity and hygroscopicity were the two significant factors that described the difference in the conductivity of exchangeable anions. Therefore, the properties and quality of the PBI membranes could be modified by selecting suitable anionic molecules in the ionic liquid compounds.
Figure 11 shows the schematic diagram for the preparation of PBI composite membranes in the research by Compañ et al. [69]. The most significant finding from the literature was that the activation energy (Eact) related to conductivity was dependent on the types of anions and obeyed the trend Eact(NTF2) < Eact(Cl) < Eact(BF4) < Eact(NCS). Consequently, anion selection was showed to be vital to improve the conductivity of the prepared ionic liquid PA-PBI membranes. The NCS anion showed the highest conductivity due to the presence of N and C atoms, which initiated more hydrogen bonds with the PA and PBI monomers, thus creating extra pathways for proton transfer. The activation energies for all membranes range from 65 to 84 kJ/mol, suggesting that the ionic conduction in these membranes primarily occurs through the vehicle-type mechanism.
Hyperbranched cross-linker ImOPBI-x membrane had outstanding oxidation stability, higher proton conductivity and satisfactory tensile strength, which meets the requirements for HT-PEMFC applications. This research also proved the ability of imidazolium groups to adsorb more phosphoric acid as well as stabilizing these molecules via delocalization and formation of hydrogen bonds due to the conjugated ring structure. Fuel cells equipped with this membrane showed a power density of 638 mW/cm2 at 160 °C and had good durability under a hydrogen/oxygen atmosphere, proving their ability in anhydrous proton exchange membrane applications [117]. Figure 12 shows the schematic steps for the preparation of hyperbranched cross-linker ImOPBI-x membrane, and after adsorption of phosphoric acid, imidazolium group ionic liquids hold more phosphoric acid in the polymer structure, thus increasing the membrane conductivity.
Mishra et al. [122] explained the mechanism of proton conductivity of doped phosphoric acid in the presence of ionic liquid compounds. They synthesized an AB-PBI membrane with 1-(3-trimethoxysilylpropyl)-2-methylimidazolium tetrafluoroborate to act as ionic liquid compounds. The formation of hydrogen bonds between hydrogen molecules from phosphoric acid with nitrogen molecules at PBI and ionic liquid structures helps the movement of proton along the polymer chain, thus initiating the conductivity of the polymer. Even though less phosphoric acid is adsorbed in the polymer chain, protonation still occurred as the presence of ionic liquid compounds held the phosphoric acid molecules via the formation of hydrogen bonds. Liu et al. [123] also suggested the possible proton transfer in the ionic liquid-based PBI membrane. The H-N bond from the ammonium cation of [dema][TfO] might interact with the C=N bond of PBI and proton transfer from H-N bond to C=N to C=N amine. A high content of [dema][TfO] provides more free protic ionic liquid, which acts as a proton conductor to improve proton transfer. Figure 13 depicts the mechanism of proton conduction in the phosphoric acid-doped AB-PBI membrane with the presence of ionic liquid compounds.
Among the protic ionic liquids, the most commonly explored is N,N-diethyl-N-methylammonium triflate ([dema][TfO]) (Figure 14) [124]. Sen et al. [125] and Niu et al. [71] concluded that PBI-[dema][TfO] membrane had higher ionic conductivity and stability compared to the other ionic liquids, performing better and generating 144 and 62 mW/cm2 at 125 °C and 250 °C [71]. Pant et al. [124] conducted molecular-dynamic (MD) simulations of [dema][TfO] doped PBI, finding higher membrane conductivity with increased doped ionic liquids, both for simulations and experimental findings. This may be due to the formation of well-developed ionic channels and the presence of free mobile ions. Table 4 lists the conductivity of different protic ionic liquids at their operating temperature.
Based on the review of several studies, it can be concluded that the presence of the ionic liquids in the phosphoric acid-doped PBI membrane has significant effects, especially on the membrane conductivity, mechanical strength and stability. Table 5 lists the ionic liquid compounds that have been applied in PA-PBI membrane in the literature and outcomes.

5. Future Prospects

To date, the development of fuel cell technology is wide open, and this niche area currently attracts various researchers, including industrial players, to study effective fuel cells. The sustainable fuel cell must include the reliable cost of manufacture, which is related to the production, storage, transportation and distribution of hydrogen, and maximum output generated [151]. The effective design of PBI-based membranes doped either with organic or inorganic materials must be finalized; thus, higher performance and durability of the fuel cell will be achieved.
Most importantly, if phosphoric acid is required in the PBI-based membrane, the leaking problem must be resolved. Currently, the application of polymeric ionic liquids in the fabrication of PA-PBI membrane seems a promising solution, as polymeric ionic liquid forms additional networks in the membrane, reinforcing the membrane structure, thus enhancing the mechanical property of the PEM. However, there are several considerations for upscaling, such as low-cost production, simple process, optimum stoichiometric ratios, reaction conditions, purification, recovery and production of films [152].

6. Conclusions

The addition of ionic liquid compounds may significantly improve the performance of PA-PBI membranes for fuel cell application. Aside from proton donors, the ionic liquid compounds can also act as retention agents to prevent the leaching of phosphoric acid in the PA-PBI membrane. Moreover, the mechanical and thermal stability and proton conductivity of the PA-PBI membrane can be modified via the selection of the anion or cation compounds in the ionic liquid structure to achieve the perfect and workable polymer electrolyte for the HT-PEMFC application. The proton conductivity of ionic liquid PA-PBI is initiated by the formation of a hydrogen bond network between ionic liquid molecules and nitrogen atoms in the PBI membrane, thus increasing the specific conductivity and preventing loss of voltage in the fuel cells. Future work in this area is necessary to explore other preparation routes or different ionic liquid compounds, especially poly(ionic liquids), as this area is still relatively new for fuel cell applications. Strengthening the backbone of polymers is needed to increase the mechanical strength of fuel cells, thus prolonging the durability and making the fuel cell industry more economical.

Author Contributions

Conceptualization, L.K.S. (Leong Kok Seng), M.S.M. and L.K.S. (Loh Kee Shyuan); writing—original draft preparation, L.K.S. (Leong Kok Seng); writing—review and editing, L.K.S. (Leong Kok Seng), M.S.M. and L.K.S. (Loh Kee Shyuan); visualization, L.K.S. (Leong Kok Seng), M.S.M. and L.K.S. (Loh Kee Shyuan); supervision, M.S.M. and L.K.S. (Loh Kee Shyuan); project administration, M.S.M.; funding acquisition, M.S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Universiti Kebangsaan Malaysia, grant number DPK-2020-009, and grant number PP-FKAB-2021.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This research was supported by Universiti Kebangsaan Malaysia under research code DPK-2020-009 and PP-FKAB-2021.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ye, Y.S.; Rick, J.; Hwang, B.J. Water soluble polymers as proton exchange membranes for fuel cells. Polymer 2012, 4, 913–963. [Google Scholar] [CrossRef] [Green Version]
  2. Mohammad, N.; Mohamad, A.B.; Kadhum, A.A.H.; Loh, K.S. A review on synthesis and characterization of solid acid materials for fuel cell applications. J. Power Sources 2016, 322, 77–92. [Google Scholar] [CrossRef]
  3. Özdemir, Y.; Ozkan, N.; Devrim, Y. Fabrication and characterization of cross-linked polybenzimidazole based membranes for high temperature PEM fuel cells. Electrochim. Acta 2017, 245, 1–13. [Google Scholar] [CrossRef]
  4. Yusoff, Y.N.; Loh, K.S.; Wong, W.Y.; Daud, W.R.W.; Lee, T.K. Sulfonated graphene oxide as an inorganic filler in promoting the properties of a polybenzimidazole membrane as a high temperature proton exchange membrane. Int. J. Hydrogen Energy 2020, 45, 27510–27526. [Google Scholar] [CrossRef]
  5. Ying, Y.; Kamarudin, S.; Masdar, M. Silica-related membranes in fuel cell applications: An overview. Int. J. Hydrogen Energy 2018, 43, 16068–16084. [Google Scholar] [CrossRef]
  6. Yusoff, Y.; Samad, S.; Loh, K.S.; Lee, T.K. Structural and morphological study of sulfonated graphene oxide prepared with different precursors. J. Kejuruter. 2018, 2, 65–71. [Google Scholar]
  7. Mohammad, N.; Mohamad, A.B.; Kadhum, A.A.H.; Loh, K.S. Effect of silica on the thermal behaviour and ionic conductivity of mixed salt solid acid composites. J. Alloys Compd. 2017, 690, 896–902. [Google Scholar] [CrossRef]
  8. Escorihuela, J.; Olvera-Mancilla, J.; Alexandrova, L.; del Castillo, L.F.; Compañ, V. Recent progress in the development of composite membranes based on polybenzimidazole for high temperature proton exchange membrane (PEM) fuel cell applications. Polymers 2020, 12, 1861. [Google Scholar] [CrossRef]
  9. Husaini, T.; Malaysia, U.K.; Sulong, A.B.; Muhammad, S.; Utara, U.S. Effect of temperature on the mechanical performance of highly conductive composites for HT-PEMFC application. J. Kejuruter. 2018, 1, 25–30. [Google Scholar] [CrossRef]
  10. Goh, J.; Rahim, A.A.; Masdar, M.; Shyuan, L. Enhanced performance of polymer electrolyte membranes via modification with ionic liquids for fuel cell applications. Membranes 2021, 11, 395. [Google Scholar] [CrossRef]
  11. Wang, G.; Yu, Y.; Liu, H.; Gong, C.; Wen, S.; Wang, X.; Tu, Z. Progress on design and development of polymer electrolyte membrane fuel cell systems for vehicle applications: A review. Fuel Process. Technol. 2018, 179, 203–228. [Google Scholar] [CrossRef]
  12. Kim, D.J.; Choi, D.H.; Park, C.H.; Nam, S.Y. Characterization of the sulfonated PEEK/sulfonated nanoparticles composite membrane for the fuel cell application. Int. J. Hydrogen Energy 2016, 41, 5793–5802. [Google Scholar] [CrossRef]
  13. Krastev, V.; Falcucci, G.; Jannelli, E.; Minutillo, M.; Cozzolino, R. 3D CFD modeling and experimental characterization of HT PEM fuel cells at different anode gas compositions. Int. J. Hydrogen Energy 2014, 39, 21663–21672. [Google Scholar] [CrossRef]
  14. Asensio, J.A.; Sanchez, E.; Gomez-Romero, P. Proton-conducting membranes based on benzimidazole polymers for high-temperature PEM fuel cells. A chemical quest. Chem. Soc. Rev. 2010, 39, 3210–3239. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, H.; Shen, P.K. Recent development of polymer electrolyte membranes for fuel cells. Chem. Rev. 2012, 112, 2780–2832. [Google Scholar] [CrossRef]
  16. Ribeirinha, P.; Alves, I.; Vázquez, F.V.; Schuller, G.; Boaventura, M.; Mendes, A. Heat integration of methanol steam reformer with a high-temperature polymeric electrolyte membrane fuel cell. Energy 2017, 120, 468–477. [Google Scholar] [CrossRef]
  17. Savage, J.; Voth, G.A. Persistent subdiffusive proton transport in perfluorosulfonic acid membranes. J. Phys. Chem. Lett. 2014, 5, 3037–3042. [Google Scholar] [CrossRef] [PubMed]
  18. Bose, S.; Kuila, T.; Nguyen, T.X.H.; Kim, N.H.; Lau, K.-T.; Lee, J.H. Polymer membranes for high temperature proton exchange membrane fuel cell: Recent advances and challenges. Prog. Polym. Sci. 2011, 36, 813–843. [Google Scholar] [CrossRef]
  19. Alberti, G.; Narducci, R.; di Vona, M.L.; Giancola, S. More on Nafion conductivity decay at temperatures higher than 80 °C: Preparation and first characterization of in-plane oriented layered morphologies. Ind. Eng. Chem. Res. 2013, 52, 10418–10424. [Google Scholar] [CrossRef]
  20. Subianto, S. Recent advances in polybenzimidazole/phosphoric acid membranes for high-temperature fuel cells. Polym. Int. 2014, 63, 1134–1144. [Google Scholar] [CrossRef]
  21. Li, Q.; He, R.; Jensen, J.O.; Bjerrum, N.J. PBI-based polymer membranes for high temperature fuel cells—Preparation, characterization and fuel cell demonstration. Fuel Cells 2004, 4, 147–159. [Google Scholar] [CrossRef]
  22. Colomban, P. Proton Conductors: Solids, Membranes, and Gels: Materials and Devices; Cambridge University Press: Cambridge, UK, 1992. [Google Scholar]
  23. Baozhong, X.; Savadogo, O. The effect of acid doping on the conductivity of polybenzimidazole (PBI). J. New Mater. Electrochem. Syst. 1999, 2, 95–101. [Google Scholar]
  24. Baozhong, X.; Savadogo, O. Hydrogen/oxygen polymer electrolyte membrane fuel cells (PEMFCs) based on alkaline-doped polybenzimidazole (PBI). Electrochem. Commun. 2000, 2, 697–702. [Google Scholar]
  25. Ma, Y.-L.; Wainright, J.S.; Litt, M.H.; Savinell, R.F. Conductivity of PBI membranes for high-temperature polymer electrolyte fuel cells. J. Electrochem. Soc. 2004, 151, A8–A16. [Google Scholar] [CrossRef]
  26. Jones, J.H.; Rozière, J. Recent advances in the functionalization of polybenzimidazole and polyetherketone for fuel cell applications. J. Membr. Sci. 2001, 185, 41–58. [Google Scholar] [CrossRef]
  27. Gilois, B.; Goujon, F.; Fleury, A.; Soldera, A.; Ghoufi, A. Water nano-diffusion through the Nafion fuel cell membrane. J. Membr. Sci. 2020, 602, 117958. [Google Scholar] [CrossRef]
  28. Zhang, Q.; Zhao, L.; Yang, H.; Kong, L.; Ran, F. Alkali-tolerant polymeric gel electrolyte membrane based on cross-linked carboxylated chitosan for supercapacitors. J. Membr. Sci. 2021, 629, 119083. [Google Scholar] [CrossRef]
  29. De Haro, J.C.; Tatsi, E.; Fagiolari, L.; Bonomo, M.; Barolo, C.; Turri, S.; Bella, F.; Griffini, G. Lignin-based polymer electrolyte membranes for sustainable aqueous dye-sensitized solar cells. ACS Sustain. Chem. Eng. 2021, 9, 8550–8560. [Google Scholar] [CrossRef]
  30. Abdel-Wahed, M.S.; El-Kalliny, A.S.; Badawy, M.I.; Attia, M.S.; Gad-Allah, T.A. Core double-shell MnFe2O4@rGO@TiO2 superparamagnetic photocatalyst for wastewater treatment under solar light. Chem. Eng. J. 2020, 382, 122936. [Google Scholar] [CrossRef]
  31. Amici, J.; Torchio, C.; Versaci, D.; Dessantis, D.; Marchisio, A.; Caldera, F.; Bella, F.; Francia, C.; Bodoardo, S. Nanosponge-based composite gel polymer electrolyte for safer Li-O2 batteries. Polymers 2021, 13, 1625. [Google Scholar] [CrossRef]
  32. Radzir, N.N.M.; Abu Hanifah, S.; Ahmad, A.; Hassan, N.H.; Bella, F. Effect of lithium bis (trifluoromethylsulfonyl) imide salt-doped UV-cured glycidyl methacrylate. J. Solid State Electrochem. 2015, 19, 3079–3085. [Google Scholar] [CrossRef]
  33. Wang, Z.; Dong, K.; Wang, D.; Luo, S.; Liu, Y.; Yi, T.; Wang, Q.; Zhang, Y.; Hao, A. In situ construction of multibuffer structure 3D CoSn@SnOx/CoOx@C anode material for ultralong life lithium storage. Energy Technol. 2020, 8, 1–8. [Google Scholar]
  34. Devanathan, R. Recent developments in proton exchange membranes for fuel cells. Energy Environ. Sci. 2008, 1, 101–109. [Google Scholar] [CrossRef]
  35. Wainright, J.S.; Wang, J.-T.; Weng, D.; Savinell, R.F.; Litt, M. Acid-doped polybenzimidazoles: A new polymer electrolyte. J. Electrochem. Soc. 1995, 142, L121–L123. [Google Scholar] [CrossRef]
  36. Vogel, H.; Marvel, C.S. Polybenzimidazoles, new thermally stable polymers. J. Polym. Sci. 1961, 50, 511–539. [Google Scholar] [CrossRef]
  37. Das, A.; Ghosh, P.; Ganguly, S.; Banerjee, D.; Kargupta, K. Salt-leaching technique for the synthesis of porous poly (2, 5-benzimidazole) (ABPBI) membranes for fuel cell application. J. Appl. Polym. Sci. 2018, 135, 45773. [Google Scholar] [CrossRef]
  38. Borjigin, H.; Stevens, K.A.; Liu, R.; Moon, J.D.; Shaver, A.T.; Swinnea, S.; Freeman, B.D.; Riffle, J.S.; McGrath, J.E. Synthesis and characterization of polybenzimidazoles derived from tetraaminodiphenylsulfone for high temperature gas separation membranes. Polymer 2015, 71, 135–142. [Google Scholar] [CrossRef] [Green Version]
  39. Hwang, K.; Kim, J.-H.; Kim, S.-Y.; Byun, H. Preparation of polybenzimidazole-based membranes and their potential applications in the fuel cell system. Energies 2014, 7, 1721–1732. [Google Scholar] [CrossRef]
  40. Zhu, L.; Swihart, M.T.; Lin, H. Unprecedented size-sieving ability in poly-benzimidazole doped with polyprotic acid for membrane H2/CO2 separation. Energy Environ. Sci. 2018, 11, 94–100. [Google Scholar] [CrossRef]
  41. Asensio, J.A.; Borrós, S.; Gomez-Romero, P. Proton-conducting polymers based on benzimidazoles and sulfonated benzimidazoles. J. Polym. Sci. Part A Polym. Chem. 2002, 40, 3703–3710. [Google Scholar] [CrossRef] [Green Version]
  42. Wang, L.; Pingitore, A.T.; Xie, W.; Yang, Z.; Perry, M.L.; Benicewicz, B.C. Sulfonated PBI gel membranes for redox flow batteries. J. Electrochem. Soc. 2019, 166, A1449–A1455. [Google Scholar] [CrossRef]
  43. Kumar, V.V.; Kumar, C.R.; Suresh, A.; Jayalakshmi, S.; Mudali, U.K.; Sivaraman, N. Evaluation of polybenzimidazole-based polymers for the removal of uranium, thorium and palladium from aqueous medium. R. Soc. Open Sci. 2018, 5, 171701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Penchev, H.; Zaharieva, K.; Milenova, K.; Ublekov, F.; Dimova, S.; Budurova, D.; Staneva, M.; Stambolova, I.; Sinigersky, V.; Blaskov, V. Novel meta- and AB-polybenzimidazole/zinc oxide polymer hybrid nanomaterials for photocatalytic degradation of organic dyes. Mater. Lett. 2018, 230, 187–190. [Google Scholar] [CrossRef]
  45. Chen, D.; Yan, C.; Li, X.; Liu, L.; Wu, D.; Li, X. A highly stable PBI solvent resistant nanofiltration membrane prepared by versatile and simple cross-linking process. Sep. Purif. Technol. 2019, 224, 15–22. [Google Scholar] [CrossRef]
  46. Huang, F.; Pingitore, A.T.; Benecewicz, B.C. Electrochemical hydrogen separation from reformate using high temperature polybenzimidadole (PBI) membranes: The role of chemistry. ACS Sustain. Chem. Eng. 2020, 8, 6234–6242. [Google Scholar] [CrossRef]
  47. Montes Luna, A.D.J.; López, N.C.F.; de León, G.C.; Camacho, O.P.; Miranda, C.Y.Y.; Mercado, Y.A.P. PBI/clinoptilolite mixed-matrix membranes for binary (N2/CH4) and ternary (CO2/N2/CH4) mixed gas separation. J. Appl. Polym. Sci. 2020, 138, 50155. [Google Scholar] [CrossRef]
  48. Kim, D.Y.; Ryu, S.; Kim, H.-J.; Ham, H.C.; Sohn, H.; Yoon, S.P.; Han, J.; Lim, T.-H.; Kim, J.Y.; Lee, S.W.; et al. Highly selective asymmetric polybenzimidazole-4, 4′-(hexafluoroisopropylidene) bis (benzoic acid) hollow fiber membranes for hydrogen separation. Sep. Purif. Technol. 2020, 257, 117954. [Google Scholar] [CrossRef]
  49. Huang, F.; Pingitore, A.T.; Benicewicz, B.C. High polymer content m/p-polybenzimidazole copolymer membranes for electrochemical hydrogen separation under differential pressures. J. Electrochem. Soc. 2020, 167, 063504. [Google Scholar] [CrossRef]
  50. Mekhilef, S.; Saidur, R.; Safari, A. Comparative study of different fuel cell technologies. Renew. Sustain. Energy Rev. 2012, 16, 981–989. [Google Scholar] [CrossRef]
  51. Zeis, R. Materials and characterization techniques for high-temperature polymer electrolyte membrane fuel cells. Beilstein J. Nanotechnol. 2015, 6, 68–83. [Google Scholar] [CrossRef] [Green Version]
  52. Lobato, J.; Cañnizares, P.; Rodrigo, M.A.; Linares, J.J.; Pinar, F.J. Study of the influence of amount of PGI-H3PO4 in the catalytic layer of a high temperature PEMFC. Int. J. Hydrogen Energy 2010, 35, 1347–1355. [Google Scholar] [CrossRef]
  53. Samms, S.R.; Wasmus, S.; Savinell, R.F. Thermal stability of proton conducting acid doped polybenzimidazole in simulated fuel cell environments. J. Electrochem. Soc. 1996, 143, 1225–1232. [Google Scholar] [CrossRef]
  54. Melchior, J.-P.; Majer, G.; Kreuer, K.-D. Why do proton conducting polybenzimidazole phosphoric acid membranes perform well in high-temperature PEM fuel cells? Phys. Chem. Chem. Phys. 2016, 19, 601–612. [Google Scholar] [CrossRef] [PubMed]
  55. Araya, S.S.; Zhou, F.; Liso, V.; Sahlin, S.L.; Vang, J.R.; Thomas, S.; Gao, X.; Jeppesen, C.; Kær, S.K. A comprehensive review of PBI-based high temperature PEM fuel cells. Int. J. Hydrogen Energy 2016, 41, 21310–21344. [Google Scholar] [CrossRef]
  56. Hu, Y.; Jiang, Y.; Jensen, J.O.; Cleemann, L.N.; Li, Q. Catalyst evaluation for oxygen reduction reaction in concentrated phosphoric acid at elevated temperatures. J. Power Sources 2018, 375, 77–81. [Google Scholar] [CrossRef] [Green Version]
  57. Yu, S.; Xiao, L.; Benicewicz, B. Durability studies of PBI-based high temperature PEMFCs. Fuel Cells 2008, 8, 165–174. [Google Scholar] [CrossRef]
  58. Eguizábal, A.; Lemus, J.; Pina, M. On the incorporation of protic ionic liquids imbibed in large pore zeolites to polybenzimidazole membranes for high temperature proton exchange membrane fuel cells. J. Power Sources 2013, 222, 483–492. [Google Scholar] [CrossRef] [Green Version]
  59. Van de Ven, E.; Chairuna, A.; Merle, G.; Benito, S.P.; Borneman, Z.; Nijmeijer, K. Ionic liquid doped polybenzimidazole membranes for high temperature Proton Exchange Membrane fuel cell applications. J. Power Sources 2013, 222, 202–209. [Google Scholar] [CrossRef]
  60. Heo, P.; Kajiyama, N.; Kobayashi, K.; Nagao, M.; Sano, M.; Hibino, T. Proton conduction in Sn0.95Al0.05P2O7–PBI–PTFE composite membrane. Electrochem. Solid-State Lett. 2008, 11, B91–B95. [Google Scholar] [CrossRef]
  61. Oh, S.-Y.; Yoshida, T.; Kawamura, G.; Muto, H.; Sakai, M.; Matsuda, A. Inorganic-organic composite electrolytes consisting of polybenzimidazole and Cs-subtituted heteropoly acids and their application for medium temperature fuel cells. J. Mater. Chem. 2010, 20, 6359–6366. [Google Scholar] [CrossRef]
  62. Ghosh, S.; Maity, S.; Jana, T. Polybenzimidazole/silica nanocomposites: Organic-inorganic hybrid membranes for PEM fuel cell. J. Mater. Chem. 2011, 21, 14897–14906. [Google Scholar] [CrossRef]
  63. Singha, S.; Jana, T. Influence of interfacial interactions on the properties of polybenzamidazole/clay nanocomposite electrolyte membrane. Polymer 2016, 98, 20–31. [Google Scholar] [CrossRef]
  64. Fuentes, I.; Andrio Balado, A.; Garcia Bernabe, A.; Escorihuela Fuentes, J.; Viñas, C.; Teixidor, F.; Compañ Moreno, V. Structural and dielectric properties of cobaltacarborane composite polybenzimidazole membranes as solid polymer electrolytes at high temperature. Phys. Chem. Chem. Phys. 2018, 20, 10173–10185. [Google Scholar] [CrossRef] [PubMed]
  65. Xu, C.; Wu, X.; Wang, X.; Mamlouk, M.; Scott, K. Composite membranes of polybenzimidazole and caesium-salts-of-heteropolyacids for intermediate temperature fuel cells. J. Mater. Chem. 2011, 21, 6014–6019. [Google Scholar] [CrossRef]
  66. Escorihuela, J.; Narducci, R.; Compañ, V.; Costantino, F. Proton conductivity of composite polyelectrolyte membranes with metal-organic frameworks for fuel cell applications. Adv. Mater. Interfaces 2018, 6, 1801146. [Google Scholar] [CrossRef]
  67. Yang, J.; Liu, C.; Gao, L.; Wang, J.; Xu, Y.; He, R. Novel composite membranes of triazole modified graphene oxide and polybenzimidazole for high temperature polymer electrolyte membrane fuel cell applications. RSC Adv. 2015, 5, 101049–101054. [Google Scholar] [CrossRef]
  68. Kim, J.; Kim, K.; Ko, T.; Han, J.; Lee, J.-C. Polybenzimidazole composite membranes containing imidazole functionalized grapheme oxide showing high proton conductivity and improved physicochemical properties. Int. J. Hydrogen Energy 2021, 46, 12254–12262. [Google Scholar] [CrossRef]
  69. Compañ, V.; Escorihuela, J.; Olvera, J.; Garcia-Bernabe, A.; Andrio, A. Influence of the anion on diffusivity and mobility of ionic liquids composite polybenzimidazol membranes. Electrochim. Acta 2020, 354, 136666. [Google Scholar] [CrossRef]
  70. Baik, K.D.; Seo, I.S. Metallic bipolar plate with a multi-hole structure in the rib regions for polymer electrolyte membrane fuel cells. Applied Energy 2018, 212, 333–339. [Google Scholar] [CrossRef]
  71. Niu, B.; Luo, S.; Lu, C.; Yi, W.; Liang, J.; Guo, S.; Wang, D.; Zeng, F.; Duan, S.; Liu, Y.; et al. Polybenzimidazole and ionic liquid composite membranes for high temperature polymer electrolyte fuel cells. Solid State Ionics 2021, 361, 115569. [Google Scholar] [CrossRef]
  72. Muthuraja, P.; Prakash, S.; Shanmugam, V.; Radhakrsihnan, S.; Manisankar, P. Novel perovskite structured calcium titanate-PBI composite membranes for high-temperature PEM fuel cells: Synthesis and characterizations. Int. J. Hydrogen Energy 2018, 43, 4763–4772. [Google Scholar] [CrossRef]
  73. Guo, H.; Li, Z.; Sun, P.; Pei, H.; Zhang, L.; Cui, W.; Yin, X.; Hui, H. Enhancing proton conductivity and durability of cross-linked PBI-based high-temperature PEM: Effectively doping a novel cerium triphosphonic-isocyanurate. J. Electrochem. Soc. 2021, 168, 024510. [Google Scholar] [CrossRef]
  74. Krishnan, N.N.; Lee, S.; Ghorpade, R.V.; Konovalova, A.; Jang, J.H.; Kim, H.-J.; Hang, J.; Henkensmeier, D.; Han, H. Polybenzimidazole (PBI-OO) based composite membranes using sulfophenylated TiO2 as both filler and cross-linker, and their use in the HT-PEM fuel cell. J. Membr. Sci. 2018, 560, 11–20. [Google Scholar] [CrossRef]
  75. Abouzari-Lotf, E.; Zakeri, M.; Nasef, M.M.; Miyake, M.; Mozarmnia, P.; Bazilah, N.A. Highly durable polybenzimidazole composite membranes with phosphonated grapheme oxide for high temperature polymer electrolyte membrane fuel cells. J. Power Sources 2019, 412, 238–245. [Google Scholar] [CrossRef]
  76. Budak, Y.; Devrim, Y. Micro-cogeneration application of a high-temperature PEM fuel cell stack operated with polybenzimidazole based membranes. Int. J. Hydrogen Energy 2020, 45, 35198–35207. [Google Scholar] [CrossRef]
  77. Eren, E.O.; Ozkan, N.; Devrim, Y. Polybenzimidazole-modified carbon nanotubes as a support material for platinum-based high-temperature proton exchange membrane fuel cell electrocatalysts. Int. J. Hydrogen Energy 2020, 46, 29556–29567. [Google Scholar] [CrossRef]
  78. Escorihuela, J.; Sahuquillo, O.; Garcia-Bernabé, A.; Giménez, E.; Compañ, V. Phosphoric acid doped polybenzimidazole (PBI)/zeolitic imidazole framework composite membranes with significantly enhanced proton conductivity under low humidity conditions. Nanomaterials 2018, 8, 775. [Google Scholar] [CrossRef] [Green Version]
  79. Wu, Y.; Liu, X.; Yang, F.; Zhou, L.L.; Yin, B.; Wang, P.; Wang, L. Achieving high power density and excellent durability for high temperature proton exchange membrane fuel cells based on cross-linked branched polybenzimidazole and metal organic frameworks. J. Membr. Sci. 2021, 630, 119288. [Google Scholar] [CrossRef]
  80. Zhu, X.; Zeng, Y.; Zhang, Z.; Yang, Y.; Zhai, Y.; Wang, H.; Liu, L.; Hu, J.; Li, L. A new composite of grapheme and molecularly imprinted polymer based on ionic liquids as fuctional monomer and cross-linker for electrochemical sensing 6-benzylaminopurine. Biosens. Bioelectron. 2018, 108, 38–45. [Google Scholar] [CrossRef]
  81. Wang, L.; Ni, J.; Liu, D.; Gong, C. Effects of branching structures on the properties of phosphoric acid-doped polybenzimidazole as a membrane material for high-temperature proton exchange membrane fuel cells. Int. J. Hydrogen Energy 2018, 43, 16694–16703. [Google Scholar] [CrossRef]
  82. Tian, X.; Wang, S.; Li, J.; Liu, F.; Wang, X.; Chen, H.; Ni, H.; Wang, Z. Composite membranes based on polybenzimidazole and ionic liquid functional Si–O–Si network for HT-PEMFC applications. Int. J. Hydrogen Energy 2017, 42, 21913–21921. [Google Scholar] [CrossRef]
  83. Earle, M.J.; Seddon, K.R. Ionic liquids. green solvents for the future. Pure Appl. Chem. 2000, 72, 1391–1398. [Google Scholar] [CrossRef] [Green Version]
  84. Plechkova, N.V.; Seddon, K.R. Applications of ionic liquids in the chemical industry. Chem. Soc. Rev. 2008, 37, 123–150. [Google Scholar] [CrossRef] [PubMed]
  85. Ye, Y.-S.; Rick, J.; Hwang, B.-J. Ionic liquid polymer electrolytes. J. Mater. Chem. A 2013, 1, 2719–2749. [Google Scholar] [CrossRef]
  86. Correia, D.M.; Fernandes, L.; Martins, P.; García-Astrain, C.; Costa, C.M.; Reguera, J.; Lanceros-Méndez, S. Ionic liquid–polymer composites: A new platform for multifunctional applications. Adv. Funct. Mater. 2020, 30, 1909736. [Google Scholar] [CrossRef]
  87. Ohno, H. Ionic liquids. In Fuctional Organic Liquids; Nakanishi, T., Ed.; Wiley-VCH Verlag GmbH&Co.: Weinheim, Germany, 2019; pp. 235–250. [Google Scholar]
  88. Łuczak, J.; Paszkiewicz-Gawron, M.; Krukowska, A.; Malankowska, A.; Zaleska-Medynska, A. Ionic liquids for nano- and microstructures preparation. Part 1: Properties and multifunctional role. Adv. Colloid Interface Sci. 2016, 230, 13–28. [Google Scholar] [CrossRef]
  89. Yasuda, T.; Watanabe, M. Protic ionic liquids: Fuel cell applications. MRS Bull. 2013, 38, 560–566. [Google Scholar] [CrossRef]
  90. Mai, N.L.; Ahn, K.; Koo, Y.-M. Methods for recovery of ionic liquids—A review. Process Biochem. 2014, 49, 872–881. [Google Scholar] [CrossRef]
  91. Esperança, J.; Lop, J.N.C.; Tariq, M.; Santos, L.; Magee, J.; Rebelo, L.P. Volatility of aprotic ionic liquids—A review. J. Chem. Eng. Data 2010, 55, 3–12. [Google Scholar] [CrossRef]
  92. Grøssereid, I.; Lethesh, K.C.; Venkatraman, V.; Fiksdahl, A. New dual functionalized zwitterions and ionic liquids; Synthesis and cellulose dissolution studies. J. Mol. Liq. 2019, 292, 111353. [Google Scholar] [CrossRef]
  93. Armand, M.; Endres, F.; MacFarlane, D.; Ohno, H.; Scrosati, B. Ionic-liquid materials for the electrochemical challenges of the future. Nat. Mater. 2009, 8, 621–629. [Google Scholar] [CrossRef] [PubMed]
  94. Rynkowska, E.; Fatyeyeva, K.; Kujawski, W. Application of polymer-based membranes containing ionic liquids in membrane separation processes: A critical review. Rev. Chem. Eng. 2018, 34, 341–363. [Google Scholar] [CrossRef]
  95. Vekariya, R.L. A review of ionic liquids: Applications towards catalytic organic transformations. J. Mol. Liq. 2017, 227, 44–60. [Google Scholar] [CrossRef]
  96. Wilkes, J.S.; Levisky, J.A.; Wilson, R.A.; Hussey, C.L. Dialkylimidazolium chloroaluminate melts: A new class of room-temperature ionic liquids for electrochemistry, spectroscopy and synthesis. Inorg. Chem. 1982, 21, 1263–1264. [Google Scholar] [CrossRef]
  97. Wilkes, J.S.; Zaworotko, M. Air and water stable 1-ethyl-3-methylimidazolium based ionic liquids. J. Chem. Soc. Chem. Commun. 1992, 13, 965–967. [Google Scholar] [CrossRef]
  98. Davis, J.H., Jr. Task-specific ionic liquids. Chem. Lett. 2004, 33, 1072–1077. [Google Scholar] [CrossRef]
  99. Iqbal, B.; Muhammad, N.; Jamal, A.; Ahmad, P.; Khan, Z.U.H.; Rahim, A.; Khan, A.S.; Gonfa, G.; Iqbal, J.; Rehman, I.U. An application of ionic liquid for preparation of homogeneous collagen and alginate hydrogels for skin dressing. J. Mol. Liq. 2017, 243, 720–725. [Google Scholar] [CrossRef]
  100. Chatterjee, K.; Pathak, A.D.; Lakma, A.; Sharma, C.S.; Sahu, K.K.; Singh, A.K. Synthesis, characterization and application of a non-flammable dicationic ionic liquid in lithium-ion battery as electrolyte additive. Sci. Rep. 2020, 10, 9606. [Google Scholar] [CrossRef]
  101. Song, M.-H.; Pham, T.P.T.; Yun, Y.-S. Ionic liquid-assisted cellulose coating of chitosan hydrogel beads and their application as drug carriers. Sci. Rep. 2020, 10, 13905. [Google Scholar] [CrossRef]
  102. Kamble, B.B.; Ajalkar, B.D.; Tawade, A.K.; Sharma, K.K.; Mali, S.S.; Hong, C.K.; Bathula, C.; Kadam, A.N.; Tayade, S.N. Ionic liquid assisted synthesis of h-MoO3 hollow microrods and their application for electrochemical sensing of Imidacloprid pesticide in vegetables. J. Mol. Liq. 2021, 324, 115119. [Google Scholar] [CrossRef]
  103. Sorkhi, S.E.S.; Hashemi, M.M.; Ezabadi, A. Introduction of a novel dicationic Brönsted acidic ionic liquid based on pyrazine and its application in the synthesis of xanthenediones and 3, 4-dihydropyrimidin-2 (1 H)-ones under solvent-free conditions. Res. Chem. Intermed. 2020, 46, 2229–2246. [Google Scholar] [CrossRef]
  104. Ferreira, T.A.; Flores-Aguilar, J.F.; Santos, E.M.; Rodriguez, J.A.; Ibarra, I.S. New poly (ionic liquid) based fiber for determination of oxytetracycline in milk samples by application of SPME-CE technique. Molecules 2019, 24, 430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Wang, Y.; Zhao, D.; Chen, G.; Liu, S.; Ji, N.; Ding, H.; Fu, J. Preparation of phosphotungstic acid based poly (ionic liquid) and its application to esterification of palmitic acid. Renew. Energy 2019, 133, 317–324. [Google Scholar] [CrossRef]
  106. Pillai, P.; Mandal, A. A comprehensive micro scale study of poly-ionic liquid for application in enhanced oil recovery: Synthesis, characterization and evaluation of physicochemical properties. J. Mol. Liq. 2020, 302, 112553. [Google Scholar] [CrossRef]
  107. Barik, B.; Kumar, A.; Nayak, P.S.; Achary, L.S.K.; Rout, L.; Dash, P. Ionic liquid assisted mesoporous silica-graphene oxide nancomposite synthesis and its application for removal of heavy metal ions from water. Mater. Chem. Phys. 2020, 239, 122028. [Google Scholar] [CrossRef]
  108. Skorikova, G.; Rauber, D.; Aili, D.; Martin, S.; Li, Q.; Henkensmeier, D.; Hempelmann, R. Protic ionic liquids immobilized in phosphoric acid-doped polybenzimidazole matrix enable polymer electrolyte fuel cell operation at 200 °C. J. Membr. Sci. 2020, 608, 118188. [Google Scholar] [CrossRef]
  109. Da Trindade, L.G.; Zanchet, L.; Martins, P.C.; Borba, K.M.; Santos, R.D.; Paiva, R.D.S.; Vermeersch, L.A.; Ticianelli, E.A.; de Souza, M.O.; Martini, E.M. The influence of ionic liquids cation on the properties of sulfonated poly (ether ether ketone)/polybenzimidazole blends applied in PEMFC. Polymer 2019, 179, 121723. [Google Scholar] [CrossRef]
  110. Rajabi, Z.; Javanbakht, M.; Hooshyari, K.; Badiei, A.; Adibi, M. High temperature composite membranes based on polybenzimidazole and dendrimer amine functionalized SBA-15 mesoporous silica for fuel cells. New J. Chem. 2020, 44, 5001–5018. [Google Scholar] [CrossRef]
  111. Shirota, H.; Mandai, T.; Fukazawa, H.; Kato, T. Comparison between dicationic and monocationic ionic liquids: Liquid density, thermal properties, surface tension, and shear viscosity. J. Chem. Eng. Data 2011, 56, 2453–2459. [Google Scholar] [CrossRef]
  112. Mishra, A.K.; Kuila, T.; Kim, D.-Y.; Kim, N.H.; Lee, J.H. Protic ionic liquid-functionalized mesoporous silica-based hybrid membranes for proton exchange membrane fuel cells. J. Mater. Chem. 2012, 22, 24366–24372. [Google Scholar] [CrossRef]
  113. Liu, F.; Wang, S.; Chen, H.; Li, J.; Wang, X.; Mao, T.; Wang, Z. The impact of poly (ionic liquid) on the phosphoric acid stability of polybenzimidazole-base HT-PEMs. Renew. Energy 2021, 163, 1692–1700. [Google Scholar] [CrossRef]
  114. Atabaki, F.; Yousefi, M.H.; Abdolmaleki, A.; Kalvandi, M. Poly (3, 4-ethylenedioxythiophene): Poly (styrenesulfonic acid) (PEDOT:PSS) conductivity enhancement through addition of imidazolium-ionic liquid derivatives. Polym. Technol. Eng. 2015, 54, 1009–1016. [Google Scholar] [CrossRef]
  115. Whitten, P.G.; Nealon, D.; Saricilar, S.Z.; Wallace, G.G. Ionic liquid solvated polymer networks for strectchable electronics. Polym. Plast. Technol. Eng. 2015, 54, 310–314. [Google Scholar] [CrossRef]
  116. Liu, F.; Wang, S.; Chen, H.; Li, J.; Tian, X.; Wang, X.; Mao, T.; Xu, J.; Wang, Z. Cross-linkable polymeric ionic liquid improve phosphoric acid retention and long-term conductivity stability in polybenzimidazole based PEMs. ACS Sustain. Chem. Eng. 2018, 6, 16352–16362. [Google Scholar] [CrossRef]
  117. Gao, C.; Hu, M.; Wang, L.; Wang, L. Synthesis and properties of phosphoric-acid-doped polybenzimidazole with hyperbranched cross-linkers decorated with imidazolium groups as high-temperature proton exchange membranes. Polymers 2020, 12, 515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Wang, X.; Wang, S.; Liu, C.; Li, J.; Liu, F.; Tian, X.; Chen, H.; Mao, T.; Xu, J.; Wang, Z. Cage-like cross-linked membranes with excellent ionic liquid retention and elevated proton conductivity for HT-PEMFCs. Electrochim. Acta 2018, 283, 691–698. [Google Scholar] [CrossRef]
  119. Zhang, N.; Wang, B.; Zhao, C.; Wang, S.; Zhang, Y.; Bu, F.; Cui, Y.; Li, X.; Na, H. Quaternized poly (ether ether ketone) s doped with phosphoric acid for high-temperature polymer electrolyte membrane fuel cells. J. Mater. Chem. A 2014, 2, 13996–14003. [Google Scholar] [CrossRef]
  120. Ye, H.; Huang, J.; Xu, J.J.; . Kodiweera, N.K.A.C.; Jayakody, J.R.P.; Greenbaum, S.G. New membranes based on ionic liquids for PEM fuel cells at elevated temperatures. J. Power Sources 2008, 178, 651–660. [Google Scholar] [CrossRef]
  121. Lin, B.; Qiao, G.; Chu, F.; Zhang, S.; Yuan, N.; Ding, J. Phosphoric acid doped hydrophobic ionic liquid-based composite membranes for anhydrous proton exchange membrane application. RSC Adv. 2017, 7, 1056–1061. [Google Scholar] [CrossRef] [Green Version]
  122. Mishra, A.K.; Kim, N.H.; Lee, J.H. Effects of ionic liquid-functionalized mesoporous silica on the proton conductivity of acid-doped poly (2, 5-benzimidazole) composite membranes for high-temperature fuel cells. J. Membr. Sci. 2014, 449, 136–145. [Google Scholar] [CrossRef]
  123. Liu, S.; Zhou, L.; Wang, P.; Zhang, F.; Yu, S.; Shao, Z.; Yi, B. Ionic-liquid-based proton conducting membranes for anhydrous H2/Cl2 fuel-cell applications. ACS Appl. Mater. Interfaces 2014, 6, 3195–3200. [Google Scholar] [CrossRef]
  124. Pant, R.; Sengupta, S.; Lyulin, A.V.; Venkatnathan, A. Computational investigation of a protic ionic liquid doped poly-benzimidazole fuel cell electrolyte. J. Mol. Liq. 2020, 314, 113686. [Google Scholar] [CrossRef]
  125. Sen, S.; Goodwin, S.E.; Barbará, P.V.; Rance, G.A.; Wales, D.; Cameron, J.M.; Sans, V.; Mamlouk, M.; Scott, K.; Walsh, D.A. Gel-Polymer electrolytes based on poly(ionic liquid)/ionic liquid networks. ACS Appl. Polym. Mater. 2021, 3, 200–208. [Google Scholar] [CrossRef]
  126. Anouti, M.; Caillon-Caravanier, M.; Dridi, Y.; Galiano, H.; Lemordant, D. Synthesis and characterization of new pyrrolidinium based protic ionic liquids. good and superionic liquids. J. Phys. Chem. B 2008, 112, 13335–13343. [Google Scholar] [CrossRef] [PubMed]
  127. Susan, A.B.H.; Noda, A.; Mitsushima, S.; Watanabe, M. Brønsted acid–base ionic liquids and their use as new materials for anhydrous proton conductors. Chem. Commun. 2003, 938–939. [Google Scholar] [CrossRef] [PubMed]
  128. Tang, B.; Gondosiswanto, R.; Hibbert, D.B.; Zhao, C. Critical assessment of superbase-derived protic ionic liquids as electrolytes for electrochemical applications. Electrochim. Acta 2019, 298, 413–420. [Google Scholar] [CrossRef]
  129. Nakamoto, H.; Watanabe, M. Brønsted acid–base ionic liquids for fuel cell electrolytes. Chem. Commun. 2007, 24, 2539–2541. [Google Scholar] [CrossRef]
  130. Miran, M.S.; Yasuda, T.; Susan, A.B.H.; Dokko, K.; Watanabe, M. Binary protic ionic liquid mixtures as a proton conductor: High fuel cell reaction activity and facile proton transport. J. Phys. Chem. C 2014, 118, 27631–27639. [Google Scholar] [CrossRef]
  131. Luo, J.; Conrad, O.; Vankelecom, I.F.J. Physicochemical properties of phosphonium-based and ammonium-based protic ionic liquids. J. Mater. Chem. 2012, 22, 20574–20579. [Google Scholar] [CrossRef] [Green Version]
  132. Nakamoto, H.; Noda, A.; Hayamizu, K.; Hayashi, S.; Hamaguchi, H.-O.; Watanabe, M. Proton-conducting properties of a brønsted acid−base ionic liquid and ionic melts consisting of bi s(trifluoromethanesulfonyl)imide and benzimidazole for fuel cell electrolytes. J. Phys. Chem. C 2007, 111, 1541–1548. [Google Scholar] [CrossRef]
  133. Yoshizawa-Fujita, M.; Byrne, N.; Forsyth, M.; Macfarlane, D.R.; Ohno, H. Proton transport properties in zwitterion blends with brønsted acids. J. Phys. Chem. B 2010, 114, 16373–16380. [Google Scholar] [CrossRef]
  134. Brigouleix, C.; Anouti, M.; Jacquemin, J.; Caillon-Caravanier, M.; Galiano, H.; Lemordant, D. Physicochemical characterization of morpholinium cation based protic ionic liquids used as electrolytes. J. Phys. Chem. B 2010, 114, 1757–1766. [Google Scholar] [CrossRef]
  135. Yoshizawa, M.; Ogihara, W.; Ohno, H. Design of new ionic liquids by neutralization of imidazole derivatives with imide-type acids. Electrochem. Solid-State Lett. 2001, 4, E25–E27. [Google Scholar] [CrossRef]
  136. Fernicola, A.; Panero, S.; Scrosati, B.; Tamada, M.; Ohno, H. New types of brönsted acid–base ionic liquids-based membranes for applications in PEMFCs. ChemPhysChem 2007, 8, 1103–1107. [Google Scholar] [CrossRef] [PubMed]
  137. Luo, J.; Hu, J.; Saak, W.; Beckhaus, R.; Wittstock, G.; Vankelecom, I.F.J.; Agert, C.; Conrad, O. Protic ionic liquid and ionic melts prepared from methanesulfonic acid and 1H-1,2,4-triazole as high temperature PEMFC electrolytes. J. Mater. Chem. 2011, 21, 10426–10436. [Google Scholar] [CrossRef] [Green Version]
  138. Xiang, J.; Chen, R.; Wu, F.; Li, L.; Chen, S.; Zou, Q. Physicochemical properties of new amide-based protic ionic liquids and their use as materials for anhydrous proton conductors. Electrochim. Acta 2011, 56, 7503–7509. [Google Scholar] [CrossRef]
  139. Lalia, B.S.; Sekhon, S. Polymer electrolytes containing ionic liquids with acidic counteranion (DMRImH2PO4, R = ethyl, butyl and octyl). Chem. Phys. Lett. 2006, 425, 294–300. [Google Scholar] [CrossRef]
  140. Che, Q.; Sun, B.; He, R. Preparation and characterization of new anhydrous, conducting membranes based on composites of ionic liquid trifluoroacetic propylamine and polymers of sulfonated poly (ether ether) ketone or polyvinylidenefluoride. Electrochim. Acta 2008, 53, 4428–4434. [Google Scholar] [CrossRef]
  141. Langevin, D.; Nguyen, Q.T.; Marais, S.; Karademir, S.; Sanchez, J.-Y.; Iojoiu, C.; Martinez, M.; Mercier, R.; Judeinstein, P.; Chappey, C. High-temperature ionic-conducting material: Advanced structure and improved performance. J. Phys. Chem. C 2013, 117, 15552–15561. [Google Scholar] [CrossRef]
  142. Gao, J.; Liu, J.; Li, W.L.B. Proton exchange membrane fuel cell working at elevated temperature with ionic liquid as electrolyte. Int. J. Electrochem. Sci. 2011, 6, 6115–6122. [Google Scholar]
  143. Rogalsky, S.; Bardeau, J.-F.; Makhno, S.; Babkina, N.; Tarasyuk, O.; Cherniavska, T.; Orlovska, I.; Kozyrovska, N.; Brovko, O. New proton conducting membrane based on bacterial cellulose/polyaniline nanocomposite film impregnated with guanidinium-based ionic liquid. Polymer 2018, 142, 183–195. [Google Scholar] [CrossRef]
  144. Guo, J.; Wang, A.; Ji, W.; Zhang, T.; Tang, H.; Zhang, H. Protic ionic liquid-grafted polybenzimidazole as proton conducting catalyst binder for high-temperature proton exchange membrane fuel cells. Polym. Test. 2021, 96, 107066. [Google Scholar] [CrossRef]
  145. Koyilapu, R.; Singha, S.; Kutcherlapati, S.; Jana, T. Grafting of vinylimidazolium-type poly (ionic liquid) on silica nanoparticle through RAFT polymerization for constructing nanocomposite based PEM. Polymer 2020, 195, 122458. [Google Scholar] [CrossRef]
  146. Lin, J.; Korte, C. PBI-type polymers and acidic proton conducting ionic liquids—Conductivity and molecular interactions. Fuel Cells 2020, 20, 461–468. [Google Scholar] [CrossRef]
  147. Nag, A.; Ali, M.A.; Singh, A.; Vedarajan, R.; Matsumi, N.; Kaneko, T. N-boronated polybenzimidazole for composite electrolyte design of highly ion conducting pseudo solid-state ion gel electrolytes with a high Li-transference number. J. Mater. Chem. A 2019, 7, 4459–4468. [Google Scholar] [CrossRef]
  148. Liu, F.; Wang, S.; Li, J.; Tian, X.; Wang, X.; Chen, H.; Wang, Z. Polybenzimidazole/ionic-liquid-functional silica composite membranes with improved proton conductivity for high temperature proton exchange membrane fuel cells. J. Membr. Sci. 2017, 541, 492–499. [Google Scholar] [CrossRef]
  149. Kallem, P.; Eguizabal, A.; Mallada, R.; Pina, M.P. Constructing straight polyionic liquid microchannels for fast anhydrous proton transport. ACS Appl. Mater. Interfaces 2016, 8, 35377–35389. [Google Scholar] [CrossRef]
  150. Hooshyari, K.; Javanbakht, M.; Adibi, M. Novel composite membranes based on dicationic ionic liquid and polybenzimidazole mixtures as strategy for enhancing thermal and electrochemical properties of proton exchange membrane fuel cells applications at high temperature. Int. J. Hydrogen Energy 2016, 41, 10870–10883. [Google Scholar] [CrossRef]
  151. Wang, S.; Jiang, S.P. Prospects of fuel cell technologies. Natl. Sci. Rev. 2017, 4, 163–166. [Google Scholar] [CrossRef]
  152. Raja, R.R.S.; Rashmi, W.; Khalid, W.; Wong, W.Y.; Priyanka, J. Recent progress in the development of aromatic polymer-based proton exchange membranes for fuel cell applications. Polymers 2020, 12, 1061. [Google Scholar]
Figure 1. Schematic flow and chemical equation of a polymer electrolyte membrane fuel cell [8].
Figure 1. Schematic flow and chemical equation of a polymer electrolyte membrane fuel cell [8].
Membranes 11 00728 g001
Figure 2. Structure of the Nafion membrane. F1 and F2 are the labels of fluorine atoms in the backbone and branching [27].
Figure 2. Structure of the Nafion membrane. F1 and F2 are the labels of fluorine atoms in the backbone and branching [27].
Membranes 11 00728 g002
Figure 3. Chemical structure of (a) monomers for the most applied PBI polymers [8]. (b) Synthesis of m-PBI [3] and (c) AB-PBI polymers [41]. (d) Synthesis of the linear and cross-linking sulfuric acid-OPBI polymer [39].
Figure 3. Chemical structure of (a) monomers for the most applied PBI polymers [8]. (b) Synthesis of m-PBI [3] and (c) AB-PBI polymers [41]. (d) Synthesis of the linear and cross-linking sulfuric acid-OPBI polymer [39].
Membranes 11 00728 g003
Figure 4. Possible proton transfer pathway for the sulfuric acid-PBI polymer [40].
Figure 4. Possible proton transfer pathway for the sulfuric acid-PBI polymer [40].
Membranes 11 00728 g004
Figure 5. Mechanism of proton transfer in the phosphoric acid-doped PBI membrane. (a) Chemical structure of PBI, (b) PA protonated PBI with no free acid molecules, (c) proton transfer along with acid-benzimidazole acid, (d) proton transfer along acid-acid and (e) proton transfer along acid-H2O [51].
Figure 5. Mechanism of proton transfer in the phosphoric acid-doped PBI membrane. (a) Chemical structure of PBI, (b) PA protonated PBI with no free acid molecules, (c) proton transfer along with acid-benzimidazole acid, (d) proton transfer along acid-acid and (e) proton transfer along acid-H2O [51].
Membranes 11 00728 g005
Figure 6. (a) Class of ionic liquid compounds and their specific applications [84]; (b) preparation of aprotic and protic ionic liquids [89]; (c) synthesis of zwitterion ionic liquids [92].
Figure 6. (a) Class of ionic liquid compounds and their specific applications [84]; (b) preparation of aprotic and protic ionic liquids [89]; (c) synthesis of zwitterion ionic liquids [92].
Membranes 11 00728 g006
Figure 7. Chemical structure of the most common cations and anions widely used for ionic liquid compounds [93,94].
Figure 7. Chemical structure of the most common cations and anions widely used for ionic liquid compounds [93,94].
Membranes 11 00728 g007
Figure 8. Types of protic ionic liquids [9].
Figure 8. Types of protic ionic liquids [9].
Membranes 11 00728 g008
Figure 9. (a) Schematic diagram for the preparation of the 6FPBI membrane, (b) proton transfer mechanism, (c) proton conductivity of all membranes with 151–171% PA uptake, (d) PA uptake at 170 °C and (e) comparison of proton conductivity [113].
Figure 9. (a) Schematic diagram for the preparation of the 6FPBI membrane, (b) proton transfer mechanism, (c) proton conductivity of all membranes with 151–171% PA uptake, (d) PA uptake at 170 °C and (e) comparison of proton conductivity [113].
Membranes 11 00728 g009
Figure 10. Schematic preparation of 6FPBI-cPIL membrane with (a) preparation of cross-linkable polymeric ionic liquid via (i) synthesis of [ViBuIm]Cl, (ii) anion exchange reaction to form [ViBuIm][TFSI] and (iii) free radical polymerization of [ViBuIm][TFSI] with allyl glycidyl ether to form cPIL. (b) Preparation of 6FPBI-cPIL membrane via a solution casting method [116].
Figure 10. Schematic preparation of 6FPBI-cPIL membrane with (a) preparation of cross-linkable polymeric ionic liquid via (i) synthesis of [ViBuIm]Cl, (ii) anion exchange reaction to form [ViBuIm][TFSI] and (iii) free radical polymerization of [ViBuIm][TFSI] with allyl glycidyl ether to form cPIL. (b) Preparation of 6FPBI-cPIL membrane via a solution casting method [116].
Membranes 11 00728 g010
Figure 11. (a) Schematic preparation of PBI membrane containing BMIM ionic liquid compounds with different anions, (b) conductivity test, (c) FT-IR results, (d) TGA results, (e) stress–strain curves and (f) Fenton’s test [69].
Figure 11. (a) Schematic preparation of PBI membrane containing BMIM ionic liquid compounds with different anions, (b) conductivity test, (c) FT-IR results, (d) TGA results, (e) stress–strain curves and (f) Fenton’s test [69].
Membranes 11 00728 g011
Figure 12. (a) Preparation of hyperbranched cross-linker ImOPBI-x, (b) FT-IR results, (c) phosphoric acid retention ability and (d) proton conductivity [117].
Figure 12. (a) Preparation of hyperbranched cross-linker ImOPBI-x, (b) FT-IR results, (c) phosphoric acid retention ability and (d) proton conductivity [117].
Membranes 11 00728 g012
Figure 13. Mechanism of proton conduction in phosphoric acid-doped AB-PBI membrane with ionic liquid compound suggested by (a) Mishra et al. [122] and (b) Liu et al. [123].
Figure 13. Mechanism of proton conduction in phosphoric acid-doped AB-PBI membrane with ionic liquid compound suggested by (a) Mishra et al. [122] and (b) Liu et al. [123].
Membranes 11 00728 g013aMembranes 11 00728 g013b
Figure 14. Chemical structure of N,N-diethyl-N-methylammonium triflate ([dema][TfO]) [124].
Figure 14. Chemical structure of N,N-diethyl-N-methylammonium triflate ([dema][TfO]) [124].
Membranes 11 00728 g014
Table 1. Preparation methods for PBI polymers and their applications.
Table 1. Preparation methods for PBI polymers and their applications.
PBI PolymerMonomerMethodApplicationRemarksRef.
Sulfonated polybenzimidazole3,3′,4,4′-tetraaminobiphenyl (TAB)Polymerization Redox flow batteriesSulfonated polybenzimidazole (s-PBI) gel membrane showed high conductivity (>240 mS/cm) and low degradation during in-cell testing[42]
m and p-PBI3,3′-diaminobenzidinePolymerizationHeavy metal absorbentp-PBI polymer indicated better performance compared to m-PBI[43]
m and AB-PBI/ZnO3,4,-diaminobenzoicm-PBI/ZnO—doping ZnO in DMAc solution with m-PBI-based powder
AB-PBI/ZnO—in situ polymerization of 3,4-diaminobenzoic with zinc nitrate.
Photocatalytic degradation of organic dyesm-PBI/ZnO performed better as a photocatalyst compared to AB-PBI/ZnO[44]
PBI3,3′-diaminobenzidineCross-linked polymerization of 4,4-dicarboxydiphenyl ether with 3,3′-diaminobenzidine in phosphorus pentaoxide/methanesulfonic acidNanofiltration membranesThe cross-linked PBI membranes exhibited more than 99% retention of Rose Bengal (RB) dye in DMF solvent[45]
m and p-PBI3,3′,4,4′-tetraaminobiphenyl (TAB)Polymerization of TAB with polyphosphoric acidElectrochemical hydrogen separationPBI membranes were used as polymer electrolytes for the EHS device
Prepared PBI membranes operate the EHS device even in high CO, producing 99.6% purity of hydrogen products with very high power efficiencies (>72%)
[46]
PBI-mixed matrix membranes3,3′-diaminobenzidine (DAB)Polymerization of DAB with polyphosphoric acid, followed by a casting process to dope different percentages of zeoliteGas separationPBI-MMMs more favorable for separation of CH4[47]
Asymmetric PHI-HFA hollow fiber4,4′-(hexafluoroisopropylidene)bis(benzoic acid)Dry-jet wet spinningHydrogen-selective membranePrepared PBI-HFA had higher permselectivities of H2/N2 and H2/CO.
Higher hydrogen flow rates were recorded compared to other gas
[48]
m/p-PBI copolymer and p-PBI homopolymer3,3′,4,4′-tetraaminobiphenyl (TAB)PPA sol-gel processElectrochemical hydrogen separation The proton conductivity and mechanical properties of the polymer depend on the final membrane composition
m/p-PBI copolymer exhibited higher creep resistance compared to homopolymer p-PBI
m/p-PBI copolymer showed long-term durability and cell recoverability
[49]
Table 2. Several examples of fillers used in the preparation of PBI polymers.
Table 2. Several examples of fillers used in the preparation of PBI polymers.
FillerOutcomesRef.
CaTiO3
  • Prepared membranes with a higher content of nano CaTiO3 show higher conductivity and good oxidative stability
  • 15% nanoCTO-PBI have 32.7 mS/cm conductivity, while 5% nanoCTO-PBI indicated conductivity of 20.2 mS/cm
  • The power density and current density of 10% nanoCTO-PBI membrane at 0.6 V and 160 °C are approximately 251.4 mW/cm2 and 419 mA/cm2
[72]
Cerium triphosphonic-isocyanurate
(Ce-TOPT)
  • Addition of Ce-TOPT proton conductor to overcome acid leaching problems
  • TOPT contains three –PO3H2 preventing water-insolubility of the membrane
  • Ce-TOPT/PBI showed good mechanical properties, proton conductivity, durability and membrane selectivity
  • Conductivity of the c-mPBI/Ce-TOPT(50) reaches 0.0885 S/cm for 50% RH, 0.125 S/cm for 100% RH and 0.0363 S/cm in anhydrous conditions
  • Proton conductivity loss about 4.6% after 48 h water-washing
[73]
Sulfophenylated TiO2
  • Metal oxide acts as filler and cross-linker
  • Ionic cross-linked system changes to covalently cross-linked system via thermal curing
  • 6c-sTiO2-PBI-OO (6 wt.% TiO2) showed the highest uptake of phosphoric acid (392 wt.%) and proton conductivity of 98 mS/cm at 160 °C
  • In fuel cell applications, a power density of 356 mW/cm2 obtained by the PBI membrane with filler, 76% higher compared to non-filler PBI membrane
[74]
Phosphonated graphene oxide
  • 76.4 × 10−3 S/cm proton conductivity is achieved at 140 °C under anhydrous conditions
  • Conductivity is more stable with the addition of PGO to the membrane
  • Strong correlation between PGO content and stability of acid content
  • In fuel cell applications, a power density more than 359 mW/cm2 at 120 °C under anhydrous conditions, 75% more than non-PGO PBI membrane
[75]
Graphene oxide
  • High power density is obtained from GO/PBI, about 546 W compared to non-GO (468 W)
  • Hydrophilic structure of GO reduced acid stripping in the membrane, improving proton conductivity
[76]
Imidazole grapheme oxide (ImGO) and grapheme oxide
  • Addition of ImGO improved physicochemical properties and higher proton conductivity
  • Addition of 0.5 wt.% ImGO enhanced tensile strength (219.2 MPa) compared to 0.5 wt.% GO (215.5 MPa) and pure PBI membrane (181.0 MPa)
  • 77.52 mS/cm of proton conductivity is obtained by ImGO/PBI
  • ImGO provides an additional effective proton transfer pathway
[68]
Multiwall carbon nanotubes (MWCNTs)
  • Pt-PBI/MWCNT shows more durability compared to Pt/C and Pt/MWCNT after the 1000th cycle of voltammetry
  • The peak current and power density of Pt-PBI/MWCNT are lower than commercial grade Pt/C, caused by the nanotube-polymer blocking the framework catalytic areas
  • Power density of Pt-PBI/MWCNT slightly increased at elevated temperatures, reaching 47 mW/cm2 at 180 °C
  • Addition of MWCNTs improved the durability of the Pt catalyst
[77]
Zeolitic imidazolate framework
  • Proton conductivity increased with increasing temperature
  • A mixture of ZIF-67 and ZIF-8 showed higher proton conductivity, about 9.2 × 10−2 S/cm, indicating a synergistic effect on proton conductivity
[78]
UiO-66 metal-organic framework
  • Introduction of UiO-66 metal-organic framework constructed channels for proton transfer
  • Increased UiO-66 caused the tendency for decreasing phosphoric acid loading and swelling ratios
  • UiO-66 also increased mechanical properties, long-term stability and enhanced PA retention
  • CBOPBI-40% UiO-66 achieved proton conductivity about 0.1 S/cm and 607 mW/cm2 power density at 160 °C with gas humidification
[79]
Table 3. Some reports on the usage of ionic liquids.
Table 3. Some reports on the usage of ionic liquids.
Ionic LiquidsApplicationRemarksRef.
1-methylimidazoliumBiopolymer solvent for preparation of collagen-alginate hydrogelsIonic liquid showed a decent potential for the preparation of collagen and alginate hydrogels[99]
1,1′-(5,14-dioxo-4,6,13,15-tetraazaoctadecane-1,18-diyl) bis(3-(sec-butyl)-1H-imidazol-3-ium) bis((trifluoromethyl)-sulfonyl) imideElectrolyte additive in lithium-ion batteryA novel dicationic room temperature ionic liquid showed a remarkable potential to subsitute conventional organic carbonate electrolyte mixture
Prepared ionic liquid was safer to use at high operation temperature with no degradation, enhanced battery life, good cycling performance and Coulombic efficiency with better discharge capacities
[100]
1-ethyl-3-methylimidazolium acetateSolventIonic liquid was employed as a solvent to dissolve chitosan before coating the surface of the chitosan hydrogel beads
Simple but effective method for cellulose coating compared to other organic solvents
[101]
1-butyl-3-methylimidazolium bromideCo-solvent for preparation of h-MoO3Ionic liquid is significant for the development of hollow rod-shaped morphology h-MoO3[102]
[SO3H-Pyrazine-SO3H] ClCatalystIonic liquid was prepared accordingly to apply as a catalyst for preparation of xanthenediones and 3,4-dihydropyrimidin-2(1H)-ones under solvent-free conditions
Several advantages were achieved, including simplicity in preparing and handling the catalyst genenrality, easy workup procedure, high yields, short reaction times, catalyst can be reused and solvent-free conditions
[103]
1-allyl-3-methylimidazolium chlorideAdsorbent for determination of oxytetracycline in milk sample A simple, effective, sensitive and environmentally friendly method for determination of oxytetracycline in milk sample via SPME-CE[104]
1-(4-sulfonate)-butyl-3-vinylimidazolium Catalyst for esterification pre-treatmentA task-specific zwitterion monomer was synthesized for production of polyzwitterion support for phosphotungstic acid grafting
Phosphotungstic acid was able to immobilize in the polymer support through chemical effects, and catalytic performance is superior due to reusability of the catalyst
[105]
1-hexadecyl-3-vinylimidazolium bromideChemical agent for oil recoverySynthesized polyionic liquid (PIL) showed good salt tolerance behavior, thermal stability and wettability alteration ability
Core flooding with PIL enhanced >30% of oil recovery after water flooding
[106]
1-butyl-3-methylimidazolium chlorideGreen solvent and porogenIonic liquid assisted the development of π-π stacking and Van Der Waals interaction toward agglomeration of the grapheme oxide sheet
Ionic liquid medium also acted as a porogen to create higher surface area of composite with better active site
[107]
Table 4. List of conductivity for different protic ionic liquids at their operating temperature.
Table 4. List of conductivity for different protic ionic liquids at their operating temperature.
Protic Ionic LiquidsConductivity (mS/cm)Temperature (°C)Ref.
Pyrrolidinium nitrate50.125[126]
Pyrrolidinium hydrogen sulfate6.825[126]
Pyrrolidinium formate32.925[126]
Pyrrolidinium acetate5.925[126]
Pyrrolidinium trifluoroacetate16.425[126]
Pyrrolidinium octanoate0.825[126]
Pyrrolidinium bis(trifluoromethanesulfonyl)amide39.6130[127]
7-methyl-1,5,7-triazabicyclo[4.4.0]dec-5-ene bis(trifluoromethanesulfonyl)imide1.5430[128]
Diethylmethylammonium trifluoromethanesulfonate43120[129]
Diethylmethylammonium hydrogen sulfate1.1030[130]
Diethylmethylammonium bis(trifluoromethanesulfonyl)amide7.4030[130]
Trioctylammonium triflate0.030325[131]
Benzimidazolium bis(trifluoromethanesulfonyl)imide8.3140[132]
3-(1-butyl-1H-imidazol-3-ium-3-yl)propane-1- sulfonate 1,1,1-Trifluoro-N-(trifluoromethylsulfonyl) methanesulfoneamide1100[133]
Morpholinium formate9.9260[134]
N-methylmorpholinium formate16.7760[134]
N-ethylmorpholinium formate12.1760[134]
Methylimidazolium bis(trifluoromethanesulfonyl)imide7.2325[135]
1-methyl-pyrazole N,N- bis(trifluoromethanesulfonyl)imide1290[136]
1H-1,2,4-triazole/methanesulfonic acid149200[137]
Isobutyramide trifluoromethanesulfonate32.6150[138]
2,3-dimethyl-1-ethylimidazolium dihydrogenphosphate70120[139]
Trifluoroacetic propylamine30180[140]
Triethylammonium triflate31130[141]
1-ethyl-3-methylimidazolium hydrogen sulfate1685[142]
N-butylguanidinium tetrafluoroborate180180[143]
Table 5. List of ionic liquid compounds used in Pa-PBI membranes.
Table 5. List of ionic liquid compounds used in Pa-PBI membranes.
Ionic Liquid TypeOutcomesConductivity Ref.
2-bromo-N,N-dimetylethanamine
  • Ionic liquid compound was protonated using trifluoromethanesulfonic acid and phosphoric acid
  • Application of prepared membrane as a catalyst binder increased coverage and desorption kinetics of oxygenated species on the catalyst surface, thus improving electrode reaction kinetics and catalytic activity of Pt/C catalyst.
-[144]
Diethylmethylammonium trifluoromethanesulfonate ([dema][TfO])
1-ethyl-3-methylimidazolium trifluoromethanesulfonate ([emim][TfO])
1-methylimidazolium bis(trifluoronethane sulfonyl)imide ([C1Im][NTf2])
1-(2-Hydroxyethyl)-3-methylimidazolium bis(trifluoromethane sulfonyl)imide (HOemim][ NTf2])
  • PBI-[dema][TfO] showed better proton conductivity compared to other PBI-ionic liquid composites
  • Proton can transfer from the H-N bond at the ammonium cation to C=N to C=N to amine
  • Free amine and diethylamine continuously accept proton at cathode and transport along the hydrogen bond in the PBI-[dema][TfO] membrane
  • Low activation energy of proton conduction for [dema][TfO] is also a significant factor for higher proton conductivity
108.9 mS/cm at 250 °C[71]
Poly(vinylimidazolium) bromide (PVImBr)
  • Better interfacial properties, greater tensile strength, storage modulus, acid loading, proton conductivity and low acid leaching
0.25 S/cm at 160 °C[145]
2-sulfoethylmethylammonium triflate [2-Sema][TfO]
  • Highly Brønsted-acidic ionic liquid assisted proton transport mechanism
  • NMR spectra helps to investigate proton exchange through interaction between polar groups and water, proving the formation of hydrogen bonds in the polymer chain
10 mS/cm at 100 °C[146]
1-butyl-3-metylimidazolium bis(trifluoromethane sulfonyl)imide [BMIm][TFSI]
  • The ionic conductivity of polymer increased with increasing ionic liquid percentage
  • Prepared composites showed thermal stability in the range of 250–300 °C, with only 10% of weight loss when the temperature was higher than 350 °C
  • LSV experiment indicated potential 4.85–5.45 V, suitable for high-energy battery application
8.8 × 10−3 S/cm at 55 °C[147]
1-butyl-3-methylimidazolium dihydrogen phosphate (BMI-DHPH)
  • BMI-DHPH enhanced phosphoric acid absorption, thus increasing proton conductivity
  • Cage-liked cross-linked polymer strengthened the mechanical
  • Properties, which meet the approval level of tensile strength for HT-PEMFC application and improve ionic liquid retention
0.133 S/cm at 160 °C[118]
1-metylimidazole trimethoxysilan
  • Ionic-liquid-functional silica enhanced the membrane performance
  • Prepared membrane had better mechanical properties, higher proton conductivity due to the high ability to absorb phosphoric acid
0.106 S/cm at 170 °C[148]
1-(3-trimethoxysilylpropyl)-3-methylimidazolium chloride
  • Hydrolysis of the ionic liquid forms a Si-O-Si network, improving the level of phosphoric acid doping and proton conductivity
  • Si-O-Si network also improved the mechanical strength, chemical, and thermal stability
0.061 S/cm at 180 °C[82]
Poly[1-(3H-imidazolium)ethylene] bis(trifluoromethanesulfonyl)imide
  • Polymeric ionic liquids play a significant role in enhancing mechanical strength and proton transfer
50 mS/cm at 200 °C[149]
1,6-di(3-methylimidazolium)hexane bis (hexafluorophosphate)
1-butyl-3-methylimidazolium hexafluorophosphate
  • Introduction of dicationic ionic liquid enhanced the performance of the fuel cell
  • Dicationic ionic liquid also increased membrane conductivity as it provides a workable ionic network for proton transfer
81 mS/cm at 180 °C[150]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Seng, L.K.; Masdar, M.S.; Shyuan, L.K. Ionic Liquid in Phosphoric Acid-Doped Polybenzimidazole (PA-PBI) as Electrolyte Membranes for PEM Fuel Cells: A Review. Membranes 2021, 11, 728. https://doi.org/10.3390/membranes11100728

AMA Style

Seng LK, Masdar MS, Shyuan LK. Ionic Liquid in Phosphoric Acid-Doped Polybenzimidazole (PA-PBI) as Electrolyte Membranes for PEM Fuel Cells: A Review. Membranes. 2021; 11(10):728. https://doi.org/10.3390/membranes11100728

Chicago/Turabian Style

Seng, Leong Kok, Mohd Shahbudin Masdar, and Loh Kee Shyuan. 2021. "Ionic Liquid in Phosphoric Acid-Doped Polybenzimidazole (PA-PBI) as Electrolyte Membranes for PEM Fuel Cells: A Review" Membranes 11, no. 10: 728. https://doi.org/10.3390/membranes11100728

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop